ArticlePDF Available

Searching for promising sources of grain protectors in extracts from Neotropical Annonaceae

Authors:

Abstract and Figures

To investigate potential sources of novel grain protector compounds against Sitophilus zeamais (Coleoptera: Curculionidae), which is an important insect pest of stored cereals, this study evaluated the bioactivity of ethanolic extracts (66) prepared from 29 species belonging to 11 different genera of Neotropical Annonaceae. A screening assay demonstrated that the most pronounced bioactive effects on S. zeamais were caused by ethanolic extracts from Annona montana, A. mucosa, A. muricata, and A. sylvatica seeds, causing the death of all weevils exposed, almost complete inhibition of the F1 progeny and a drastic reduction in grain losses. Furthermore, the ethanolic extracts obtained from the leaves of A. montana, A. mucosa, A. muricata, and Duguetia lanceolata, especially A. montana and A. mucosa, demonstrated significant bioactive effects on the studied variables; however, the activity levels were less pronounced than in the seed extracts, and the response was dependent on the concentration used. This study is the first to report the activity of secondary metabolites from D. lanceolata on insects as well as the action of A. sylvatica on pests associated with stored grains.
Content may be subject to copyright.
© 2016
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas 15 (4): 215 - 232
ISSN 0717 7917
www.blacpma.usach.cl
Artículo Original | Original Article
215
Searching for promising sources of grain protectors in extracts from
Neotropical Annonaceae
[Búsqueda de fuentes prometedoras de protectores de granos em extractos de Anonaceas Neotropicales]
Leandro do Prado Ribeiro1, José Djair Vendramim1, Gabriel Luiz Padoan Gonçalves1, Thiago Felipe Ansante1,
Eduardo Micotti da Gloria2, Jenifer de Carvalho Lopes3, Renato Mello-Silva3 & João Batista Fernandes4
1Department of Entomology and Acarology and 2Agroindustry, Food and Nutrition Department
University of São Paulo/ “Luiz de Queiroz” College of Agriculture (USP/ESALQ), Piracicaba, São Paulo State, Brazil
3Department of Botany, University of São Paulo (IB/USP), São Paulo, São Paulo State, Brazil
4Department of Chemistry, Federal University of São Carlos (UFSCar), São Carlos, São Paulo State, Brazil
Contactos | Contacts: Leandro do Prado RIBEIRO - E-mail address: lpribeiro@usp.br
Abstract: To investigate potential sources of novel grain protector compounds against Sitophilus zeamais (Coleoptera: Curculionidae),
which is an important insect pest of stored cereals, this study evaluated the bioactivity of ethanolic extracts (66) prepared from 29 species
belonging to 11 different genera of Neotropical Annonaceae. A screening assay demonstrated that the most pronounced bioactive effects on
S. zeamais were caused by ethanolic extracts from Annona montana, A. mucosa, A. muricata, and A. sylvatica seeds, causing the death of all
weevils exposed, almost complete inhibition of the F1 progeny and a drastic reduction in grain losses. Furthermore, the ethanolic extracts
obtained from the leaves of A. montana, A. mucosa, A. muricata, and Duguetia lanceolata, especially A. montana and A. mucosa,
demonstrated significant bioactive effects on the studied variables; however, the activity levels were less pronounced than in the seed
extracts, and the response was dependent on the concentration used. This study is the first to report the activity of secondary metabolites
from D. lanceolata on insects as well as the action of A. sylvatica on pests associated with stored grains.
Keywords: Allelochemicals, acetogenins, bioactivity, Sitophilus zeamais, stored cereals.
Resumen: Para investigar las posibles fuentes de nuevos compuestos protectores de granos contra Sitophilus zeamais (Coleoptera:
Curculionidae), una importante plaga de los cereales almacenados, este estudio evaluó la bioactividad de los extractos etanólicos (66)
preparados a partir de 29 especies pertenecientes a 11 géneros distintos de Anonaceas Neotropicales. Un ensayo de selección demostró que
los efectos bioactivos más relevantes sobre S. zeamais fueron causados por los extractos etanólicos de las semillas de Annona montana, de
A. mucosa, de A. muricata y de A. sylvatica, que causaron la muerte de todos los gorgojos expuestos, la inhibición parcial de la progenie F1
y una drástica reducción de las pérdidas de grano. Además, los extractos etanólicos obtenidos de las hojas de A. montana, de A. mucosa, de
A. muricata y de Duguetia lanceolata, especialmente de A. montana y de A. mucosa, demostraron efectos bioactivos significativos sobre las
variables estudiadas. Sin embargo, los niveles de bioactividad fueron menores en comparación con los extractos de semillas, y la respuesta
fue dependiente de la concentración utilizada. Este estudio es el primer relato sobre la actividad de los metabolitos secundarios de D.
lanceolata sobre insectos, así como la acción de A. sylvatica sobre plagas asociadas a los granos almacenados..
Palabras clave: Aleloquímicos, acetogeninas, bioactividad, Sitophilus zeamais, cereales almacenados.
Recibido | Received: May 15, 2014
Aceptado | Accepted: May 19, 2015
Aceptado en versión corregida | Accepted in revised form: February 21, 2016
Publicado en línea | Published online: July 30, 2016
Declaración de intereses | Declaration of interests: the São Paulo Research Foundation (FAPESP, grant number 2010/52638-0) and the National Science and Technology Institute
for Biorational Control of Pest Insects (INCT-CBIP, grant number 573742/2008-1) for the financial support..
Este artículo puede ser citado como / This article must be cited as: LP Ribeiro, JD Vendramim, GLP Gonçalves, TF Ansante, EM da Gloria, JC Lopes, R Mello-Silva & JB
Fernandes. 2016. Searching for promising sources of grain protectors in extracts from Neotropical Annonaceae Bol Latinoam Caribe Plant Med Aromat 15 (4): 215 232.
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/216
INTRODUCTION
The large diversity of secondary metabolites in plants
(allelochemicals) originates from a long evolutionary
process that relies on the relationships between the
plants and their competitors (natural enemies). Plants
develop these compounds mainly as a defense
mechanism (Wink, 2003). The study of
allelochemicals has not only increased the knowledge
of the processes involved in the interactions between
plants and other factors in the environment but has
also led to the discovery of important bioactive
molecules of great interest for humankind (Ramesha
et al., 2011). Among other functions, these bioactive
molecules are used as model-prototypes for the
development of new drugs (Miller, 2011) and new
products for the protection of agricultural crops and
stored commodities (Cantrell et al., 2012).
The structural and functional diversity of
allelochemicals is a key factor for the survival and
evolutionary success of plant species inhabiting an
environment with an abundance of natural enemies.
Therefore, the tropical flora, with its unique
biodiversity, is a promising natural reservoir of
bioactive substances (Valli et al., 2012). In this
context, Brazil exhibits enormous potential for the
development of novel active substances based on
natural products because the country has the highest
plant genetic diversity in the world, with more than
55,000 catalogued species (Simões & Schenkel,
2002). However, to date, this potential has not been
well exploited.
Among the botanical families that occur in
the Neotropical regions, Annonaceae is the main
family of the order Magnoliales (APG III, 2009) and
is one of the most specious families of angiosperms
comprising 135 genera and approximately 2,500
species (Chatrou et al., 2004). Annonaceae exhibits a
pantropical distribution with 40 genera and 900
species in the Neotropical region. In Brazil, this
family is represented by 29 genera, of which 1 are
endemic, and 386 species, and a large proportion of
this richness is found in the Amazon Rain Forest and
Atlantic Forest (Maas et al., 2013).
Despite the lack of studies, a large number of
diverse chemical compounds present in the different
structures of Annonaceae plants have been isolated.
Alkaloids, acetogenins, diterpenes and flavonoids are
the main chemical groups in extracts from the bark,
branches, leaves, fruits and seeds of Annonaceae
(Lebouef et al., 1982; Chang et al., 1998; Kotkar et
al., 2001). Among these classes, acetogenins are
conspicuous because of the vast array of biological
activities they exhibit. Acetogenins are a series of
natural products (C-35/C-37) derived from long-
chain fatty acids (C-32/C-34) combined with a 2-
propanol unit (Alali et al., 1999).
Our previous studies (Ribeiro, 2010; Ribeiro
et al., 2013) demonstrated a promising grain-
protectant effect in seed extracts from two species of
Annona, which are characterized by a complex
mixture of acetogenins and alkaloids. This
observation motivated additional biomonitoring
investigations in other Annonaceae species in order
to explore more comprehensively the richness of
allelochemicals in this plant family and the species
diversity of the Brazilian flora. These studies
prompted our research program to search for
allelochemicals with activities against pest species of
stored grains, which is an essential component of
current stored grain integrated pest management
programs (IPM).
This study evaluated the bioactivity of
ethanolic extracts (66) of different structures from 29
Annonaceae species (7.5% of all Brazilian species)
belonging to 11 different genera (Anaxagorea,
Annona, Duguetia, Ephedranthus, Guatteria,
Hornschuchia, Oxandra, Porcelia, Pseudoxandra,
Unonopsis, and Xylopia) against Sitophilus zeamais
Motschulsky (Coleoptera: Curculionidae), which is
an important pest of stored cereals under tropical
conditions. In addition, the fungicidal and
antiaflatoxigenic activities of the promising extracts
were evaluated in vitro against the isolate CCT7638
of Aspergillus flavus Link (Ascomycota: Eurotiales:
Trichocomaceae), a producer of aflatoxin B1, in order
to better characterize the potential of these extracts as
grain protectors.
MATERIAL AND METHODS
Species sampling and plant extract preparation
The collection data for the plant species used in the
study, which were obtained from different locations in
the south and southeast regions of Brazil, are shown in
Table 1. In total, 29 species of Annonaceae belonging
to 11 genera were collected and were investigated.
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/217
For the extract preparation, the plant structures
were collected and were dehydrated in an oven at 40º
C for 48 to 72 hours. Subsequently, the materials
were separately milled in a knife mill to obtain a
powder of each plant structure, which were stored
separately in sealed glass containers until use. The
organic extracts were obtained by maceration in
ethanol solvent (1:5, w v-1). For this step, the plant
powder was maintained in the solvent for 3 days after
which it was filtered through filter paper. The
remaining residual cake was placed back in the
ethanol solvent, and this process was repeated 4
times. The solvent remaining in the filtered solution
was eliminated in a rotary evaporator at 50º C and -
600 mm Hg pressure. After the solvent was
evaporated in the airflow chamber, the extraction
yield of each structure for all species was determined.
Bioassays
The bioassays were performed in a climate-controlled
room at 25 ± C, 60 ± 10% relative humidity, a
photoperiod of 14 hours and a mean luminosity of
200 lux. Whole corn grains were used as the substrate
for the assays. The corn grains were manually
selected from the hybrid AG 1051 (yellow dent,
semi-hard) from crops developed without
insecticides. The experimental design used for the
tests was completely randomized.
A microatomizer coupled to a pneumatic
pump and adjusted to a pressure of 0.5 kgf cm-2 with
a spray volume of 30 L t-1 was used for the
application of the treatments. After spraying, the
grains/extract mixture was manually placed in 2-liter
plastic bags, which were lightly shaken for 1 minute.
Screening for the identification of promising
extracts
To identify extracts with bioactivity against S.
zeamais, bioassays were performed to verify the
insecticidal activity and sublethal effects, which were
assessed by evaluating the number of insects emerged
(F1 progeny) and the damage to the treated samples.
As a large number of extracts was obtained, they
were divided into 5 groups in order to be tested. The
groups were established according to the similarity of
collection dates and plant structures available for
each species. The extracts were assayed at
concentration of 3,000 mg kg-1 (mg of extract kg-1
of corn), which were defined based on previous
studies (Ribeiro, 2010).
Evaluation of insecticidal activity
For this bioassay, corn samples (10 g, in Petri dishes
measuring 6-cm in diameter × 2-cm high) were
treated separately with the respective extracts. The
growth substrate treated with the solvent
[acetone:methanol solution (1:1, v v-1)] was used as
the control. Preliminary assays were performed to
evaluate the possible effects of the solution used for
the resuspension of the extracts on S. zeamais. Next,
each Petri dish was infested with 20 adult S. zeamais
(aged 10 to 20 days) from both sexes, and 10
replicates per treatment were performed. The adult
survival was evaluated on day 10 after infestation.
The insect was considered dead when its extremities
were completely distended and it exhibited no
reaction to contact with a paintbrush for 1 minute.
Evaluation of F1 progeny and damages
The same sampling units used for the insecticidal
assay were used in this bioassay. The grains were
treated with the respective extracts and were infested
with 20 adults from both sexes (aged 10 to 20 days).
After 10 days of infestation, the adults were removed,
and the sampling units were kept under the climate
conditions previously described. As before, 10
replicates per treatment were performed.
At 60 days after the initial infestation, the
number of emerged adults in each dish was counted.
The damages caused by the feeding of the S. zeamais
were determined through the visual verification of the
percentage of damaged or perforated grains in each
sample. In addition, the grain weight loss (%) was
estimated based on the equation proposed by Adams
and Schulten (1976).
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/218
Table 1
Species and structures of Annonaceae used in the study and the respective collection data.
Species
Local of collection
Data of
collection
Voucher
(Herbarium)1
Anaxagorea dolichocarpa
Sprague & Sandwith
Vale Natural Reserve, Linhares, ES,
Brazil
(19º08’04,0” S; 40º03’24,5” W;
elevation: 33 m)
17/11/2011
Lopes 361
(CVRD, ESA, SPF)
Annona acutiflora
Mart.
Vale Natural Reserve, Linhares, ES,
Brazil
(19º09’05,4” S; 40º04’02,4” W;
elevation: 70 m)
16/11/2011
Lopes 144
(CVRD, ESA, SPF)
Annona cacans
Warm.
IAC/APTA, Jundiaí, SP, Brazil
(23º06’56,6” S; 46º55’58,3” W;
elevation: 599 m)
04/03/2011
Ribeiro 17
(ESA)
Annona dolabripetala
Raddi
Botanical Garden of São Paulo, São
Paulo, SP, Brazil
(23º32’18,2” S; 46º36’44,5” W;
elevation: 745 m)
07/06/2011
Ribeiro 18
(ESA)
Annona emarginata
(Schltdl.) H.Rainer
IAC/APTA, Jundiaí, SP, Brazil
(23º06’53,4” S; 46º56’07,2” W;
elevation: 616 m)
04/03/2011
Ribeiro 16
(ESA)
Annona montana
Macfad.
ESALQ/USP Campus, Piracicaba,
SP, Brazil
(22º42’28,2” S; 47º37’59.4” W;
elevation: 537 m)
21/03/2011
Ribeiro 3
(ESA)
Annona mucosa
Jacq.
ESALQ/USP Campus, Piracicaba,
SP, Brazil
(22º42’28,5” S; 47º37’59.6” W;
elevation: 534 m)
17/03/2011
Ribeiro 4
(ESA)
Annona muricata
L.
ESALQ/USP Campus, Piracicaba,
SP, Brazil
(22º42’25,4” S; 47º37’43,9” W;
elevation: 576 m)
12/04/2011
Ribeiro 12
(ESA)
Annona reticulata
L.
ESALQ/USP Campus, Piracicaba,
SP, Brazil
(22º42’51,4” S; 47º37’38,8” W;
elevation: 548 m)
01/03/2011
Ribeiro 11
(ESA)
Annona sp. 1
São Luís Farmer, Descalvado, SP,
Brazil
(2152’58,0” S; 4740’38,0” W;
elevation: 679 m)
02/04/2011
Ribeiro 13
(ESA)
Annona sp. 2
Frutas Raras Farmer, Rio Claro, SP,
Brazil
(23º06’53,4” S; 46º56’07,2” W;
elevation: 716 m)
02/04/2011
Ribeiro 14
(ESA)
Annona sylvatica
A.St.-Hil.
Ribeiro Small Farmer, Erval Seco,
RS, Brazil
(27º25’41,8” S; 53º34’11,2” W;
elevation: 466 m)
25/04/2011
Ribeiro 10
(ESA)
Duguetia lanceolata
A.St. Hil.
ESALQ/USP Campus, Piracicaba,
SP, Brazil
(22º42’41,5” S; 47º38’0,2” W;
elevation: 556 m)
23/03/2011
Ribeiro 9
(ESA)
Ephedranthus dimerus J.C.Lopes, Chatrou
& Mello-Silva (1)
Vale Natural Reserve, Linhares, ES,
Brazil
(19º09’05,5” S; 40º04’00,1” W;
elevation: 67 m)
17/11/2011
Lopes 145
(CVRD, ESA, SPF, WAG)
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/219
Ephedranthus dimerus J.C.Lopes, Chatrou
& Mello-Silva (2)
Vale Natural Reserve, Linhares, ES,
Brazil
(19º10’19,2” S; 40º01’08,7” W;
elevation: 48 m)
17/11/2011
Lopes 362
(CVRD, ESA, SPF, WAG)
Guatteria australis
A. St.-Hil.
Vale Natural Reserve,Linhares, ES,
Brazil
(19º08’28,5” S; 40º04’05,7” W;
elevation: 18 m)
17/11/2011
Lopes 153
(CVRD, ESA, SPF)
Guatteria ferruginea
A. St.-Hil.
Santa Lúcia Biological Station, Santa
Tereza, ES, Brazil
(19º58’02,5” S; 40º32’15,5” W;
elevation: 694 m)
13/11/2011
Lopes 348
(ESA)
Guatteria sellowiana
Schltdl.
Santa Lúcia Biological Station, Santa
Tereza, ES, Brazil
(19º58’02,5” S; 40º32’15,5” W;
elevation: 694 m)
13/11/2011
Lopes 345
(ESA)
Guatteria villosissima
A. St.-Hil.
Vale Natural Reserve, Linhares, ES,
Brazil
(19º09’25,6” S; 40º31’43,6” W;
elevation: 636 m)
16/11/2011
Lopes 146
(ESA)
Hornschuchia bryotrophe Nees
Vale Natural Reserve, Linhares, ES,
Brazil
(19º07’56,4” S; 40º05’05,7” W;
elevation: 58 m)
17/11/2011
Lopes 111
(ESA)
Hornschuchia citriodora
D.M. Johnson
Vale Natural Reserve, Linhares, ES,
Brazil
(19º07’57,8” S; 40º05’05,9” W;
elevation: 48 m)
17/11/2011
Lopes 110
(ESA)
Hornschuchia myrtillus
Nees
Vale Natural Reserve, Linhares, ES,
Brazil
(19º11’13,3” S; 39º54’49,6” W;
elevation: 44 m)
17/11/2011
Lopes 364
(CVRD)
Oxandra martiana
(Schltdl.) R E. Fr.
Vale Natural Reserve, Linhares, ES,
Brazil
(19º11’30,8” S; 39º57’09,3” W;
elevation: 31 m)
17/11/2011
Lopes 363
(CVRD, ESA, SPF)
Porcelia macrocarpa
(Warm.) R.E. Fries
Botanical Garden of São Paulo, São
Paulo, SP, Brazil
(23º33’55,8” S; 46º36’08,3” W;
elevation: 752 m)
07/06/2011
Eiten 791
(SP)
Pseudoxandra spiritus-sancti
Maas
Vale Natural Reserve, Linhares, ES,
Brazil
(19º04’44,8” S; 39º53’19,5” W;
elevation: 20 m)
16/11/2011
Lopes 317
(CVRD, SPF)
Unonopsis sanctae-teresae
Maas & Westra
Goiapaba-açu Road, Santa Teresa,
ES, Brazil
(19º54’50,5” S; 40º31’59,1” W;
elevation: 840 m)
14/11/2011
Lopes 355
(ESA)
Xylopia brasiliensis
Spreng.
Augusto Ruschi Biological Reserve,
Santa Tereza, ES, Brazil
(19º54’27,1” S; 40º33’02,0” W;
elevation: 805 m)
14/11/2011
Lopes 351
(ESA)
Xylopia decorticans
D.M. Johnson & Lobão
Augusto Ruschi Biological Reserve,
Santa Tereza, ES, Brazil
(19º54’28,5” S; 40º32’57,3” W;
elevation: 807 m)
14/11/2011
Lopes 352
(ESA)
Xylopia frutescens
Aubl.
Vale Natural Reserve, Linhares, ES,
Brazil
16/11/2011
Lopes 359
(ESA)
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/220
(19º09’23,4” S; 39º59’30,3” W;
elevation: 30 m)
Xylopia laevigata
(Mart.) R.E. Fr.
Vale Natural Reserve, Linhares, ES,
Brazil
(19º05’01,2” S; 39º53’04,8” W;
elevation: 22 m)
16/11/2011
Lopes 316
(ESA)
1 CVRD (Vale Natural Reserve Herbarium, Linhares, ES, Brazil); ESA (“Luiz de Queiroz” College of Agriculture
Herbarium, Piracicaba, SP, Brazil); SP (Botanical Institut of São Paulo Herbarium, São Paulo, Brazil); SPF
(University of São Paulo Herbarium, São Paulo, SP, Brazil); WAG (Wageningen University Herbarium,
Wageningen, Netherlands).
Concentration-response curves of active extracts
The extracts demonstrating the most promising
results were bioassayed for the estimation of the LC50
and LC90, corresponding to the concentration
necessary to kill 50% and 90% of the population of
weevils, respectively. For these estimations,
preliminary tests were performed to determine the
baseline concentrations that caused the death of 95%
of the adults and a mortality rate similar to that in the
control. Based on these results, the test
concentrations (5-7 concentrations; range: 50 - 4,000
mg kg-1) were established using the formula
proposed by Finney (1971). The remaining
experimental procedures were the same as those used
in the initial screening, in which the mortality was
assessed 10 days after the infestation of the sample
units.
Estimation of average lethal time (LT50) of the
promising extracts
For each selected extract, the time required to kill
50% of the weevil population (LT50) was estimated
based on the LC90 value determined in the previous
bioassay. The same procedures described in the
screening assay were used for this bioassay; however,
the evaluation of weevil mortality was performed
every 24 hours for 10 days.
Fungicidal and antiaflatoxigenic effects of the most
promising extracts
The antifungal and antiaflatoxigenic activities
(against isolate CCT7638 of A. flavus, a producer of
AFB1) of the most promising extracts were evaluated
using a method termed poison food (Alvarez-
Castellanos et al., 2001). This technique is based on
the observation of the growth of fungal mycelium in
YES (yeast extract saccharose) culture media using
1,000 mg L-1 of the respective extracts (final
concentration) dissolved in 5 mL solvent solution
[acetone: water, (1:3, v v-1)], incorporated by manual
agitation in unfused culture media (temperature
approximately 45º C). The extract + medium (10 mL)
were added to each Petri dish (6.5-cm diameter), and
10 dishes were used per treatment. The solvent
solution (acetone: water) was included as a control,
and water was used as the negative control.
The fungus was inoculated following the
solidification of the media as follows: the central area
of the Petri dish was perforated using a Stanley knife
previously immersed in a conidia solution. After
incubation of the fungal colonies for 11 days in PDA
(potato, dextrose and agar) media, the spore
suspension was prepared by scraping the media
surface using a Drigalski spatula followed by
immersion in 50 mL of an aqueous solution (47.5 mL
of distilled water + 2.5 mL of dimethyl sulfoxide).
The amount of conidia in the solution was
standardized to contain 2 to 9 × 105 conidia mL-1,
measured using a Neubauer chamber. The inoculated
dishes were sealed with plastic film and were
incubated upside down at 25 ± C and a scotophase
of 24 hours.
The radial mycelial growth was evaluated at
48, 96, 144 and 192 hours after the fungal
inoculation. The evaluation consisted of measuring
the diameter of each colony in 2 directions at a
straight angle using a caliper. Based on the arithmetic
mean of the 2 measurements, the percentage
inhibition (P.I.) of the treatments relative to the
control was calculated using the following equation:
100/)(.. ControlTreatmentControlIP
The production of aflatoxin (B1) by isolate
CCT7638 of A. flavus grown in culture media
containing the respective treatments was assessed
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/221
using the thin-layer chromatography (TLC)
technique. The culture media from 5 Petri dishes
randomly chosen from each treatment was transferred
to a 50-mL Falcon tube using Stanley knives and
spatulas. The weight of the transferred material was
recorded, and the material was subjected to an
aflatoxin extraction process.
For the extraction of aflatoxins from the
media, 14 mL of distilled water and 18 mL of
analysis-grade methanol were added to the tube. The
content was vortexed for 1 minute. A 5-mL aliquot of
the extract was transferred to an amber vial, and the
solution volume was evaporated entirely using a
sample concentrator with airflow at 45º C. The dried
material was redissolved in 200 μL of toluene:
acetonitrile (9:1) and agitated for 30 seconds using
ultrasound. The presence and quantification of
aflatoxins in the extract were performed in aluminum
chromatoplates with silica gel. The development of
the chromatography plates was performed in vats
containing 5 mL of the elution solution composed of
ether:methanol:water (96:3:1). The presence of
aflatoxins in the samples was verified by comparison
to the AFB1 standard. The standard AFB1 solution
was prepared based on the Sigma-Aldrich standard
(Sigma AF-1), and the concentration was determined
according to methodology 971.22 found in the
American Official Analytical Chemistry (2006). For
this assay, the concentration of aflatoxins in the
samples was compared with 3-, 4- and 6-µL aliquots
of the AFB1 standard. The calculation of
contamination was performed according to the
following equation:
Where: AF= aflatoxin content (AFB1); Y =
standard concentration in μg mL-1; S = μL of the
standard toxin with fluorescence equivalent to the
sample; V = extract final volume (sample) in μL; X =
extract initial volume (sample) in μL; W = sample
weight, in grams, in the final extract.
Data analysis
Generalized linear models (GLM) (Nelder &
Wedderburn, 1972) with quasi-binomial distributions
were used for the analysis of the proportions of
mortality and damaged grains, whereas GLM with
quasi-Poisson distributions was used for the analysis
of emerged insect numbers. GLM with Gaussian
distribution was used for the analysis of A. flavus
vegetative growth and aflatoxin production. In all
cases, the goodness-of-fit was determined using a
half-normal probability plot with a simulated
envelope (Hinde & Demétrio, 1998). When a
significant difference was observed between the
treatments, multiples comparisons (Tukey’s test, P <
0.05) were performed using the glht function of the
multicomp package with adjustment of P values.
Multivariate analyses were performed to
determine the grouping of the crude extracts of
Annonaceae based on the variables analyzed in the
screening assay. The mean Euclidian distance was
used as a measurement of similarity, and the
UPGMA (unweighted pair-group average) method
was used as a clustering strategy. The relationship
between the variables analyzed was determined using
Spearman’s nonparametric analysis (P = 0.05). The
analyses were performed using the software “R”,
version 2.15.1 (R Development Core Team, 2012).
A binomial model with a complementary
log-log link function (gompit model) was used to
estimate the lethal concentrations (LC50 and LC90),
using the Probit Procedure in the software SAS
version 9.2 (SAS Institute, 2011). Finally, the mean
lethal time (LT50) was estimated using the method
proposed by Throne et al. (1995) for Probit analysis
of correlated data.
RESULTS
Selection of the promising crude extracts
Of the 66 tested extracts at 3,000 mg kg-1, the most
pronounced bioactive effects on S. zeamais were
caused by the ethanolic extracts from the A. montana,
A. mucosa, A. muricata and A. sylvatica seeds (Table
2). These extracts produced complete mortality of the
exposed weevils, almost total inhibition of the F1
progeny and a drastic reduction in damage to the
treated samples. The ethanolic extracts from the A.
montana, A. mucosa, A. muricata and D. lanceolata
leaves, especially the A. montana and A. mucosa
leaves, exhibited significant bioactive effects;
however, compared with the seed extracts, the effects
were at lower levels and exhibited a concentration-
dependent response.
 
)/()( XWYSVAF
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/222
Table 2
Mortality (mean ± standard error) on day 10, number of emerged insects (F1 progeny) and damage after 60
days of infestation with Sitophilus zeamais (Coleoptera: Curculionidae) in corn samples (10 g) treated with
ethanolic extracts from different species and/or structures of Annonaceae (3,000 mg kg-1)*.
Species
Plant
structures
Mortality
(%)1
No. of emerged
insects2
% Grains
damaged1
Grain weight losses
Total (%)3
Relative (%)4
Group A[2]
Annona cacans
Leaves
0.50±0.50 c
51.00±5.12 a
86.98±6.99 a
10.87±0.87
96.71
Branches
0.00±0.00
54.20±2.64 a
89.09±1.54 a
11.13±0.19
99.02
Seeds
14.50±2.03 c
41.50±4.11 ab
76.20±3.01 a
9.52±0.37
84.70
Annona montana
Leaves
77.50±5.73 a
8.40±3.35 cd
17.71±6.90 c
2.21±0.86
19.66
Branches
1.50±0.76 c
46.90±2.24 a
82.01±2.57 a
10.25±0.32
91.19
Seeds
100.00±0.00
0.00±0.00
0.00±0.00
0.00±0.00
0.00
Annona mucosa
Leaves
83.00±5.33 a
6.10±1.79 d
13.54±3.84 c
1.69±0.48
15.04
Branches
0.00±0.00
59.90±3.92 a
92.22±2.68 a
11.52±0.33
102.49
Seeds
100.00±0.00
0.00±0.00
0.00±0.00
0.00±0.00
0.00
Annona muricata
Leaves
34.00±6.27 b
25.90±4.61 bc
48.31±7.50 b
6.03±0.93
53.65
Branches
1.50±0.76 c
47.40±2.74 a
86.38±2.81 a
10.79±0.35
96.00
Seeds
100.00±0.00
0.00±0.00
0.00±0.00
0.00±0.00
0.00
Annona sylvatica
Leaves
19.50±4.18 b
41.40±4.32 ab
74.67±4.49 ab
9.36±0.56
83.27
Branches
0.00±0.00
46.70±2.84 a
83.39±3.30 a
10.42±0.41
92.70
Seeds
100.00±0.00
0.00±0.00
0.00±0.00
0.00±0.00
0.00
Duguetia lanceolata
Leaves
37.50±4.60 b
17.00±2.74 c
42.35±6.13 b
5.29±0.76
47.06
Branches
1.00±0.66 c
51.04±2.71 a
87.58±2.21 a
10.94±0.27
97.33
Seeds
0.50±0.50 c
51.00±2.40 a
89.23±1.93 a
11.15±0.24
99.20
Control
(acetone:methanol, 1:1
(v/v))
--
0.00±0.00
51.40±3.72 a
89.97±3.10 a
11.24±0.38
--
F
57.24
24.24
25.94
P value
< 0.0001
< 0.0001
< 0.0001
Group B[2]
Annona dolabripetala
Leaves
4.50±1.74 ab
28.50±2.70 c
61.90±3.74 d
7.73±0.46
70.98
Branches
0.00±0.00
57.80±2.07 ab
88.44±2.41 abc
11.05±0.30
101.47
Annona emarginata
Leaves
0.00±0.00
46.90±2.79 b
84.49±3.65 abc
10.56±0.45
96.97
Branches
0.00±0.00
55.1±1.64 ab
95.10±0.48 a
11.88±0.06
109.09
Annona reticulata
Leaves
5.00±2.23 ab
43.10±3.98 b
82.08±4.19 abc
10.26±0.52
94.21
Branches
3.00±1.10 ab
44.50±4.43 b
79.71±6.60 c
9.96±0.82
91.46
Annona sp.1
Leaves
0.50±0.50 b
47.30±3.26 b
81.03±3.23 bc
10.12±0.40
92.93
Branches
0.00±0.00
68.40±3.47 a
94.35±1.10 abc
11.79±0.13
108.26
Annona sp. 2
Leaves
16.50±4.15 a
27.90±3.32 c
58.57±4.56 d
7.32±0.57
67.22
Branches
0.00±0.00
54.80±2.41 ab
88.75±1.63 abc
11.09±0.20
101.84
Porcelia macrocarpa
Leaves
0.00±0.00
68.30±1.66 a
94.72±1.31 ab
11.84±0.16
108.72
Branches
0.50±0.50 b
53.20±3.00 ab
91.78±1.80 abc
11.47±0.22
105.33
Control
(acetone:methanol, 1:1
(v/v))
--
2.00±1.52 ab
51.00±2.48 b
87.17±2.64 abc
10.89±0.33
--
F
7.23
16.29
12.38
P value
<0.0001
<0.0001
<0.0001
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/223
Group C[2]
Annona acutiflora
Leaves
7.00±2.26
27.70±2.57
65.84±3.45
8.23±0.43
94.93
Branches
4.50±1.38
37.50±3.27
81.37±3.97
10.17±0.49
117.30
Guatteria australis
Leaves
6.00±2.76
32.30±3.50
74.80±3.58
9.35±0.44
107.84
Branches
6.50±1.50
32.90±3.90
72.72±4.75
9.09±0.59
104.84
Guatteria ferruginea
Leaves
9.00±2.66
25.00±3.09
62.16±3.28
7.77±0.41
89.62
Branches
8.50±2.24
30.00±1.61
74.67±3.30
9.33±0.41
107.61
Guatteria sellowiana
Leaves
7.50±3.00
33.90±4.37
73.82±6.04
9.22±0.75
106.34
Branches
2.00±0.81
35.70±2.64
78.21±2.73
9.77±0.34
112.69
Guatteria villosissima
Leaves
3.50±1.50
30.50±3.00
69.52±5.72
8.69±0.71
100.23
Branches
2.00±1.10
38.20±2.45
77.57±2.81
9.69±0.35
111.76
Unonopsis sanctae-
teresae
Leaves
4.00±1.45
32.70±2.13
74.51±3.49
9.31±0.43
107.38
Branches
3.00±1.10
28.60±2.85
67.41±5.18
8.42±0.64
97.12
Control
(acetone:methanol, 1:1
(v/v))
--
3.00±1.33
29.00±2.60
69.37±5.17
8.67±0.64
100.00
F
1.81ns
1.66 ns
1.65 ns
P value
0.0540
0.0834
0.0868
Group D[2]
Oxandra martiana
Leaves
0.00±0.00
46.80±2.96 ab
87.98±3.45
10.99±0.43
101.85
Branches
2.50±0.83
45.20±4.18 ab
85.78±3.58
10.72±0.44
99.35
Pseudoxandra spiritus-
sancti
Leaves
3.00±1.52
38.00±1.69 b
83.21±2.14
10.40±0.26
96.39
Branches
4.00±2.33
41.90±3.91ab
80.60±5.03
10.07±0.62
93.33
Xylopia brasiliensis
Leaves
0.00±0.00
48.70±2.72 ab
90.20±2.94
11.27±0.36
104.45
Branches
1.00±0.66
49.80±2.48 ab
90.01±1.74
11.25±0.21
104.26
Xylopia decorticans
Leaves
0.50±0.50
44.30±2.81 ab
86.21±2.79
10.77±0.34
99.81
Branches
0.50±0.50
41.80±2.90 ab
86.16±2.79
10.77±0.34
99.81
Xylopia frutescens
Leaves
0.00±0.00
44.20±4.76 ab
83.65±4.44
10.45±0.55
96.85
Branches
0.50±0.50
56.40±4.83 a
89.72±3.66
11.21±0.45
103.89
Xylopia laevigata
Leaves
1.00±0.66
38.50±1.63 b
81.53±2.50
10.19±0.31
94.44
Branches
0.50±0.50
45.60±2.71 ab
88.24±2.48
11.03±0.31
102.22
Control
(acetone:methanol, 1:1
(v/v))
--
0.50±0.50
41.50±2.73 ab
86.35±2.85
10.79±0.35
100.00
F
1.78 ns
2.31
0.94 ns
P value
0.8097
0.0110
0.5079
Group E[2]
Anaxagorea
dolichocarpa
Leaves
9.50±3.11 a
35.60±3.46 ab
75.23±3.50 ab
9.40±0.43
97.11
Branches
4.50±1.74 a
47.60±4.59 a
85.66±3.35 ab
10.70±0.41
110.54
Ephedranthus dimerus
1
Leaves
1.50±1.06 a
39.50±4.48 ab
77.55±6.33 ab
9.69±0.79
100.10
Branches
0.50±0.50 a
34.40±3.21 ab
79.98±4.41 ab
9.99±0.55
103.20
Ephedranthus dimerus
2
Leaves
8.50±2.98 a
32.40±4.90 ab
71.11±8.13 ab
8.88±1.01
91.74
Branches
1.00±0.66 a
45.50±2.68 a
89.06±2.52 a
11.13±0.31
114.98
Hornschuchia
bryotrophe
Leaves
7.50±2.81 a
23.80±2.36 b
60.89±5.75 b
7.61±0.71
78.62
Branches
3.00±0.81 a
38.80±4.92 ab
77.94±5.48 ab
9.74±0.68
100.62
Hornschuchia
citriodora
Leaves
8.00±3.81 a
44.00±4.48 a
85.62±4.46 ab
10.70±0.55
110.54
Branches
1.00±0.66 a
34.60±4.07 ab
74.66±4.40 ab
9.33±0.55
96.38
Hornschuchia myrtillus
Leaves
12.00±5.22 a
27.88±4.96 ab
62.60±6.92 b
7.82±0.66
80.79
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/224
Branches
2.00±1.10 a
48.80±2.96 a
87.45±2.07 a
10.93±.025
112.91
Control
(acetone:methanol, 1:1
(v/v))
--
1.00±0.66 a
37.40±5.89 ab
77.48±6.17 ab
9.68±0.77
100.00
F
3.72
3.09
2.84
P value
0.0115
0.0108
0.0190
1Means followed by different letters in the columns containing each tested group extracts indicate
significant differences between treatments (GLM with quasi-binomial distribution followed by Tukey’s post
hoc test, P<0.05); 2Means followed by different letters in the columns containing each tested group extracts
indicate significant differences between treatments (GLM with quasi-Poisson distribution followed by
Tukey’s post hoc test, P<0.05); 3Calculated using the formula proposed by Adams and Schulten (1976);
4Calculated based on the relative comparison of the treatment (extract) with its respective control;
*Applied using a spray volume of 30 L t-1; ns: Not significant (P>0.05)
The hierarchical grouping analysis using the
data from the variables analyzed in the screening
bioassays indicated the formation of 3 groups (Figure
1). The first group comprised the ethanolic extracts
from the A. montana, A. mucosa, A. muricata and A.
sylvatica seeds, as well as extracts from the A.
montana and A. mucosa leaves, which demonstrated
the most pronounced lethal and sublethal effects. The
second group comprised the ethanolic extracts from
D. lanceolata and A. muricata leaves, species that
demonstrated less pronounced bioactive effects
(lethal and sublethal). The third group encompassed
the controls and extracts that did not demonstrate
bioactivity against the targeted pest.
Independently of the concentration tested, the
adult mortality was inversely correlated with the
other tested variables [F1 progeny (r = -0.59; P <
0.0001) and % damaged grains(r = -0.58; P <
0.0001)]. Although mortality was the variable with
the highest weight in the separation between
treatments, based on the Spearman’s correlation
coefficients, one cannot discard the small
oviposition- and/or feeding-deterrent action of the
extracts from the respective Annonaceae species.
Estimation of lethal concentrations and average
lethal time of the selected extracts
The extract prepared from the A. mucosa seeds
demonstrated the lowest LC50 and LC90 values
(288.33 and 505.47 mg kg-1, respectively) (Table 3).
These values were significantly different from the
values for the other active extracts based on the
comparison of the estimated confidence intervals (P
< 0.05). In general, the extracts from the seeds of
different bioactive Annona species demonstrated the
lowest values compared with the active extracts from
leaves. On the other hand, the average lethal time
(varying between 82.06 and 94.85 hours) did not
demonstrate great differences between the treatments
(Table 4), showing a slower activity of the active
extracts, which is probably related with the
mechanisms of action of the active compounds.
Fungicidal and antiaflatoxigenic activity of the most
promising extracts
Overall, the ethanolic extracts tested (1,000 mg kg-1),
which were selected based their activity on S.
zeamais, did not significantly inhibit the vegetative
growth of the isolate CCT7638 of A. flavus and did
not affect the production of AFB1 after 192 hours of
incubation (Table 5). However, the extract from the
A. sylvatica seeds reduced the initial growth rate
(fungistatic effect) of the radial mycelial growth after
48 hours of incubation. Based on the results, the
toxicity of the solvent used for extract solubilization
(acetone) in A. flavus was verified. Although the
concentration was low (25%), caution is
recommended when using this organic solvent in
these types of bioassays.
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/225
Figure 1
Dendrogram obtained from Cluster analysis based on bioactivity similarity of ethanolic extracts from
Annonaceae on Sitophilus zeamais (Coleoptera: Curculionidae) [mean Euclidian distance as dissimilarity
measurement and UPGMA (Unweighted Pair Group Method) as a clustering strategy method]
at 3,000 mg kg-1.
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/226
Table 3
Estimation of LC50 and LC90 (in mg kg-1*) and confidence interval of ethanolic extracts from Annonaceae
for Sitophilus zeamais (Coleoptera: Curculionidae) adults after 10 days of exposure in treated maize
samples (10 g).
Species
(structures)
n 1
Slope ± SE
(valor de P)
LC50
(CI) 2
LC90
(CI) 2
χ2 (3)
d.f. 4
h.5
Annona montana
(leaves)
1,400
4.86±0.29
(P<0.0001)
1,851.00
(1,758.00 1,942.00)
3,270.00
(3,075.00 3,516.00)
3.75
4
0.94
Annona montana
(seeds)
1,200
7.09±0.42
(P<0.0001)
621.70
(557.11677.44)
942.45
(858.961,071.91)
5.55
3
1.85
Annona mucosa
(leaves)
1,400
5.74±0.46
(P<0.0001)
1,972.00
(1,847.002,080.00)
3,190.00
(3,026.003,405.00)
1.56
4
0.82
Annona mucosa
(seeds)
1,600
4.92±0.33
(P <0.0001)
288.33
(267.29307.21)
505.47
(478.58537.75)
4.71
5
0.94
Annona muricata
(seeds)
1,600
6.35±0.37
(P <0.0001)
384.94
(364.47403.75)
594.76
(569.35624.19)
4.88
5
0.98
Annona muricata
(leaves)
--
--
>3,000
--
--
--
--
Annona sylvatica
(seeds)
1,200
6.32±0.75
(P <0.0001)
554.48
(471.11617.90)
858.58
(799.49918.24)
3.07
3
0.37
Duguetia lanceolata
(leaves)
--
--
>3,000
--
--
--
--
1 n: number of tested insects; 2 CI: 95% confidence interval; 3 χ2: calculated chi-squared value;
4 d.f.: degrees of freedom; 5 h.: heterogeneity factor; * Applied using a spray volume of 30 L t-1;
-- Not determined.
Table 4
Estimation of average lethal time (LT50, in hours) and confidence interval of ethanolic extracts from
Annonaceae for Sitophilus zeamais (Coleoptera: Curculionidae) adults.
Species
(structures)
n 1
Slope ± SE
LT50 (CI) 2
χ2 (3)
d.f. 4
h.5
Annona montana
(leaves)
400
8.30±0.41
88.76
(83.3994.34)
18.19
7
2.60
Annona montana
(seeds)
400
7.08±0.30
86.67
(82.1291.08)
15.39
8
1.98
Annona mucosa
(leaves)
400
5.98±0.24
94.85
(88.89100.68 )
13.67
7
1.95
Annona mucosa
(seeds)
400
6.31±0.18
86.13
(84.0488.17)
7.60
8
0.95
Annona muricata
(seeds)
400
5.61±0.22
90.42
(84.3296.23)
19.14
8
2.39
Annona sylvatica
(seeds)
400
7.18±0.23
82.06
(78.8185.21)
16.69
8
2.09
1 n: number of tested insects; 2 CI: 95% confidence interval;
3 χ2: calculated chi-squared value; 4 d.f.: degrees of freedom;
5 h.: heterogeneity factor.
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/227
Table 5
Radial growth (mean ± standard error) of isolate CCT7638 of Aspergillus flavus colonies and production of
aflatoxin (AFB1) in YES (yeast extract saccharose) culture media containing ethanolic seed extracts from
different Annonaceae (1,000 mg L-1).
Extracts / Incubation time
Diameter of colonies (mm) 1
Production of
48 hours
P.I.
(%)2
96 hours
P.I.
(%)2
144 hours
P.I.
(%)2
192 hours
P.I.
(%)2
AFB1 (ppm mm-
2) 1
Annona montana
10.30±0.88
b
+27.16
37.90±2.02
b
+22.65
50.80±2.09
ab
+ 25.43
54.77±1.24
ab
+ 10.65
42.99±5.21
Annona mucosa
9.30±1.22
bc
+14.81
33.30±3.26
b
+ 7.77
44.20±3.56
bc
+ 9.16
53.40±1.97
ab
+ 7.88
41.91±7.32
Annona muricata
7.30±0.20 c
- 9.87
29.00±0.83
bc
- 6.15
43.00±2.05
bc
+ 6.17
53.80±1.07
ab
+ 8.69
42.23±3.94
Annona sylvatica
4.80±0.23
d
- 40.74
25.40±0.91
c
- 17.80
38.50±1.98
c
- 4.94
52.00±1.71
ab
+ 5.05
40.82±6.77
Control (acetone:water, 1:3
(v/v))
8.10±0.83
bc
--
30.90±2.71
bc
--
40.50±2.85
c
--
49.50±2.66
b
--
38.85±5.11
Negative control (water)
15.60±0.25
a
--
46.30±1.29 a
--
54.10±1.34
a
--
55.95±0.05
a
--
43.92±4.49
F
37.662
18.21
8.0269
2.4677
0.6521ns
P value
<0.0001
<0.0001
<0.0001
0.04378
0.6439
1 Means followed by different letters, in the columns, indicate significant differences between treatments
(GLM with Gaussian distribution followed by Tukey’s post hoc test, P<0.05)
2 P.I.: Percentage of inhibition;
ns: Not significant (P>0.05).
DISCUSSION
Our study provides important information regarding
promising sources of compounds to be used as grain
protectors in the preventative management of
Coleoptera pest species in stored cereals. Annonaceae
is one of the most diverse and abundant plant family
in Neotropical forests (Chatrou et al., 2004; Maas et
al., 2011a; Maas et al., 2011b) and has been shown to
exhibit secondary metabolites with great chemical
diversity and promising biological activities (Lebouef
et al., 1982; Zafra-Polo et al., 1998; Colom et al.,
2010); however, this study represents the most
comprehensive screening study performed to date.
Despite the variations in soil-climate
conditions in the sampling sites, which could have
influenced the chemical profiles of the extracts, and
the differences in the sampling effort for the different
plant structures, genus Annona seeds were identified
as the main sites of accumulation of compounds with
activity against insects. Therefore, our results are
consistent with other reports in the literature
(Leatemia & Isman, 2004; Llanos et al., 2008; Seffrin
et al., 2010; Grzybowski et al., 2013; Ribeiro et al.,
2013). Concerning the action on insects, this study is
the first to report the activity of compounds derived
from A. sylvatica (formerly Rollinia sylvatica) on
pests associated with stored grains. To date, a small
number of secondary compounds from A. sylvatica,
native to the center-south region of Brazil (Lorenzi et
al., 2005), were isolated and evaluated for their
bioactive potential. Consistent with our results,
Mikolajczak et al. (1990) demonstrated the oral
toxicity of a hexanic extract from A. sylvatica fruits
against Ostrinia nubilalis (Hübner) (Lepidoptera:
Crambidae) larvae. After successive fractionation, the
compound sylvaticin (the only acetogenin reported
from this species to date) was isolated and shown to
exhibit a series of biological activities, including the
protection of cantaloupe plants against Acalymma
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/228
vittata Barber (Coleoptera: Chrysomelidae).
Recently, Formagio et al. (2013) demonstrated the
anti-inflammatory and anti-carcinogenic properties of
essential oils of A. sylvatica leaves, which are mostly
composed of a combination of oxygenated
sesquiterpenes.
A number of reports in the literature ratify the
putative toxicological effects (acute or chronic) of
derivative compounds from the remaining species of
Annona for medically and agriculturally relevant
pests. Therefore, compounds from A. muricata
demonstrate activities against Plutella xylostella
(Linnaeus) (Lepidoptera: Plutellidae) (Trindade et al.,
2011), Bactericera cockerelli (Sulc) (Hemiptera:
Triozidae) (Flores-Davila et al., 2011), Anastrepha
ludens (Loew) (Diptera: Tephritidae) (Gonzalez-
Esquinca et al., 2012) and Aedes aegypti (Linnaeus)
(Diptera: Culicidae) (Grzybowski et al., 2013).
Acetogenins isolated from A. montana seeds
demonstrate toxicity for Oncopeltus fasciatus
(Dallas) (Hemiptera: Lygaeidae) (Colom et al., 2008)
and insecticidal and anti-feeding activities for
Spodoptera frugiperda (J.E. Smith) (Lepidoptera:
Noctuidae) larvae (Blessing et al., 2010).
Furthermore, the extract obtained through the
decoction of A. mucosa seeds has a repellent effect on
Acromyrmex octospinosus (Forel) (Hymenoptera:
Formicidae) workers (Boulogne et al., 2012).
Using S. zeamais as a model, Llanos et al.
(2008) demonstrated the efficient control of adults
and the complete inhibition of the F1 progeny using
extracts (at concentrations higher than 2,500 ppm)
from A. muricata seeds using hexane and ethyl
acetate as solvents. In previous studies, we
demonstrated that A. montana (Ribeiro, 2010) and A.
mucosa (Ribeiro et al., 2013) seed extracts in both
hexane and dichloromethane solvents caused
promising bioactive effects on S. zeamais. In contrast
to the solvents used in this study, the extractions in
the previous study were performed using organic
solvents at gradients of increasing polarity in all
cases. Given the changes in the chemical profiles of
the derivatives obtained using different techniques
and/or solvents, this difference can partially explain
the differences in bioactive concentrations observed
between the studies. Meanwhile, this study
demonstrates the possibility of extracting active
principles of these species using a “green solvent”
that is naturally biodegradable and produced from
renewable sources.
A large number of compounds of diverse
chemical natures in several structures of the genus
Annona have been isolated in a number of
phytochemical studies (Lebouef et al., 1982; Chang
et al., 1998; Kotkar et al. 2001). Among the
compounds, acetogenins stand out because of their
structural abundance and the wide array of biological
activities they exhibit, such as powerful insecticidal
and acaricidal activities (Alali et al., 1999; Colom et
al., 2008, 2010). Acetogenins are potent inhibitors of
complex I (NADH:ubiquinone oxidoreductase) of the
mitochondrial electron-transport system and of the
enzyme NADH:oxidase in the cell membrane of
target arthropods (Lewis et al., 1993). According to
Bermejo et al. (2005), the majority of acetogenins in
Annonaceae have been isolated from the seeds and
stems of the genera Annona, Anomianthus, Asimina,
Desepalum, Goniothalamus, Rollinia [now Annona
(Rainer, 2007)], Polyalthia, Porcelia, Uvaria and
Xylopia. However, the relative content of the
derivatives originating from the different genera and
the structure-activity relationship of the acetogenins
from different structures remain to be investigated.
Based on the estimated lethal time (in the
LC90 estimated for each selected extract), the slower
action of the active compounds was confirmed. The
symptomatology of the contaminated insects was
characterized by the inactivity, locomotive instability
and food avoidance, followed by the collapse,
paralysis, and slow death by respiratory insufficiency.
These signs are typical of the action of compounds
that inhibit mitochondrial respiration, such as
rotenone and piericidin (Ware and Whitacre, 2004.).
Therefore, based on these findings and previous
analysis (Ribeiro et al., 2013), it is possible to
hypothesize that the biological activity of the extracts
from A. sylvatica seeds and A. montana, A. mucosa
and A. muricata seeds and leaves are because of the
presence of acetogenins.
Despite the effects were less expressive
compared with those observed for the extracts of the
genus Annona, this study is the first to report the
activity of D. lanceolata derivatives on pest insects, a
plant species that has not been studied from a
phytochemical point of view. However, a number of
pharmacological properties, such as antiprotozoal
(Tempone et al., 2005), antinociceptive and anti-
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/229
inflammatory activities (Sousa et al., 2008), of crude
ethanolic extracts and/or the isoquinoline alkaloid-
rich fraction from D. lanceolata leaves have been
reported in the literature and corroborate the potential
observed in this study.
To better elucidate the potential protectant
effect on grains, the antifungal and antiaflatoxigenic
activities of the most promising extracts (regarding
the action on S. zeamais) were evaluated using the A.
flavus isolate. Although the fungal toxicity of the
crude extracts and acetogenins isolated from the
genus Annona species were reported previously for a
number of plant species (Dang et al., 2011) and
human pathogens (Ahmad and Sultana, 2003; Lima et
al., 2011), the extracts evaluated did not exert
pronounced effects on A. flavus. Hypothetically, the
lack of fungicidal effect could be because of
evolutionary selection in A. flavus species adapted to
coexist with these secondary metabolites.
Corroborating this hypothesis, Okwulehie and Alfred
(2010) demonstrated that A. flavus is a species that
deteriorates A. muricata fruits in Nigeria. However,
because the insects are the main agent of dispersion
of fungal spores in the grain mass, the adequate
control of pest insect species provided by the
respective extracts can decrease the incidence of A.
flavus and consequently the aflatoxin levels in the
stored grains.
Despite the preliminary nature of the data in
this study, it can be concluded that ethanolic extracts
from A. sylvatica seeds, D. lanceolata leaves and A.
montana, A. mucosa and A. muricata seeds and
leaves exert bioactive effects on S. zeamais.
Accordingly, bio-guided studies are being conducted
in order to purify, isolate and characterize the
compounds responsible for the observed bioactivity.
Additionally, it will be possible to evaluate the
potential use of these compounds as model-molecules
or biorational compounds in integrated management
programs of the stored pests. However, this study
provides a scientific basis for the rational utilization
of Annonaceae species with potential use to humans,
mainly as a homemade tool for stored grain
protection.
ACKNOWLEDGMENTS
The authors thank Dr. Sérgio Sartori from the Rare
Fruits Brazilian Association (Rio Claro, SP, Brazil);
Dr. José Bettiol Neto from the Fruit Center of the
Agronomic Institute of Campinas (IAC/APTA,
Jundiaí, SP, Brazil) and Vale Natural Reserve
(Linhares, ES, Brazil), for the help with obtaining the
Annonaceae species used in this study. We also thank
the São Paulo Research Foundation (FAPESP, grant
number 2010/52638-0) and the National Science and
Technology Institute for Biorational Control of Pest
Insects (INCT-CBIP, grant number 573742/2008-1)
for the financial support.
REFERENCES
Adams JM, Schulten GGM. 1976. Losses caused by
insects, mites, and microorganisms. In
Harris KL & Lindblad CJ: Postharvest Grain
Loss and Assessment Methods. American
Association of Cereal Chemists, St. Paul,
MN, USA.
Ahmad KF, Sultana N. 2003. Biological studies on
fruit pulp and seeds of Annona squamosa. J
Chem Soc Pak 25: 331 - 334.
Alali FQ, Liu XX, McLaughlin JL. 1999.
Annonaceous acetogenins: recent progress. J
Nat Prod 62: 504 - 540.
Alvarez-Castellanos PP, Bishop CD, Pascual-
Villalobos MJ. 2001. Antifungal activity of
the essential oil of flower heads of garland
chrysanthemum (Chrysanthemum
coronarium) against agricultural pathogens.
Phytochemistry 57: 99 - 102.
American Official Analitical Chemistry. 2006.
Official methods of analysis, Item
49.2.02, AOAC International, Rockville,
USA.
APG III (The Angiosperm Philogeny Group). 2009.
An update of the Angiosperm Phylogeny
Group classification for the orders and
families of flowering plants: APG III. Bot J
Linn Soc 161: 105 - 121.
Bermejo A, Figadère B, Zafra-Polo MC, Barrachina
I, Estornell E, Cortes D. 2005. Acetogenins
from Annonaceae: recents progress in
isolation, synthesis and mechanisms of
action. Nat Prod Rep 22: 269 - 303.
Blessing LT, Colom OA, Popich S, Neske A, Bardón
A. 2010. Antifeedant and toxic effects of
acetogenins from Annona montana on
Spodoptera frugiperda. J Pest Sci 83: 307 -
310.
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/230
Boulogne I, Germosen-Robineau L, Ozier-Lafontanie
H, Jacoby-Koaly C, Aurela L, Loranger-
Merciris G. 2012. Acromyrmex octospinosus
(Hymenoptera: Formicidae) management.
Part 1: Effects of TRAMIL’s insecticidal
plant extracts. Pest Manage Sci 68: 313 -
320.
Cantrell CL, Dayan CL, Duke SO. 2012. Natural
products as sources for new pesticides. J Nat
Prod 75: 1231 - 1242.
Chang FR, Yang PY, Lin JY, Lee KH, Wu YC. 1998.
Bioactive kaurane diterpenoids from Annona
glabra. J Nat Prod 61: 437 - 439.
Chatrou LW, Rainer H, Maas PJM. 2004.
Annonaceae. In Smith N, Mori SA,
Henderson A, Stevenson DW, Heald SV:
Flowering plants of the Neotropics, Princeton
University Press, Published in association
with The New York Botanical Garden,
Princeton, USA.
Colom OA, Barrachina I, Mingol IA, Mas MCG,
Sanz PM, Neske A, Bardon A. 2008. Toxic
effects of annonaceous acetogenins on
Oncopeltus fasciatus. J Pest Sci 81: 85-89.
Colom OA, Salvatore A, Willink E, Ordonez R, Isla
MI, Neske A, Bardon A. 2010. Insecticidal,
mutagenic and genotoxic evaluation of
annonaceous acetogenins. Nat Prod
Commun 5: 391-394.
Dang QL, Kim WK, Nguyen CM, Choi YH, Choi GJ,
Jang KS, Park MS, Lim CH, Luu NH, Kim
JC. 2011. Nematicidal and antifungal
activities of annonaceous acetogenins from
Annona squamosa against various plan
pathogens. J Agric Food Chem 59: 11160 -
11167.
Finney DJ. 1971. Probit analysis. Cambridge
University Press, Cambridge, UK.
Flores-Davila M, Gonzalez-Villegas R, Guerrero-
Rodriguez E, Mendoza-Villareal R,
Cardenas-Elizondo A, Cerna-Chavez E,
Aguirre-Uribe L. 2011. Insecticidal effect of
plant extracts on Bactericera cockerelli
(Hemiptera: Psyllidae) nymphs. Southwest
Entomol 36: 137 - 144.
Formagio ASN, Vieira MD, Dos Santos LAC,
Cardoso CAL, Foglio MA, De Carvalho JE,
Andrade-Silva M, Kassuya CAL. 2013.
Composition and evaluation of the anti-
inflammatory and anticancer activities of the
essential oil from Annona sylvatica A. St.-
Hil. J Med Food 16: 20 - 25.
Gonzalez-Esquinca AR, Luna-Cazares LM, Guzman
MAS, Chacon ID, Hernandez GL, Breceda
SF, Gerardo PM. 2012. In vitro larvicidal
evaluation of Annona muricata L., A.
diversifolia Saff. and A. lutescens Saff.
extracts against Anastrepha ludens larvae
(Diptera, Tephritidae). Interciencia 37: 284 -
289.
Grzybowski A, Tiboni M, Silva MAN, Chitolina RF,
Passos M, Fontana JD. 2013. Synergistic
larvicidal effect and morphological
alterations induced by ethanolic extracts of
Annona muricata and Piper nigrum against
the dengue fever vector Aedes aegypti. Pest
Manage Sci 69: 589 - 601.
Hinde J, Demétrio CGB. 1998. Overdispersion:
models and estimation. Comput Stat Data
Anal 27: 151 - 170.
Kotkar HM, Prashant SM, Sangeetha VGSS, Shipra
RG, Shripad MU, Maheshwari VL. 2001.
Antimicrobial and pesticidal activity of
partially puried avonoids of Annona
squamosa. Pest Manage Sci 58: 33 - 37.
Leatemia JA, Isman MB. 2004. Toxicity and
antifeedant activity of crude seed extracts of
Annona squamosa (Annonaceae) against
lepidopteran pests and natural enemies. Int J
Trop Insect Sci 24: 150 - 158.
Leboeuf M, Cave A, Bhaumik PK, Mukherjee B,
Mukherjee R. 1982. The phytochemistry of
the Annonaceae. Phytochemistry 21: 2783 -
2813.
Lewis MA, Arnason JT, Philogene BJR, Rupprecht
JK, McLaughlin JL. 1993. Inhibition of
respiration at site I by asimicin, an
insecticidal acetogenin of the pawpaw,
Asimina triloba (Annonaceae). Pestic
Biochem Physiol 45: 15 - 23.
Lima LARS, Johann S, Cisalpino PS, Pimenta LPS,
Boaventura MAD. 2011. Antifungal activity
of 9-hydroxy-folianin and sucrose octaacetate
from the seeds of Annona cornifolia A. St. -
Hil. (Annonaceae). Food Res Int 44: 2283 -
2288.
Llanos CAH, Arango DL, Giraldo MC. 2008.
Actividad insecticida de extractos de semilla
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/231
de Annona muricata (Annonaceae) sobre
Sitophilus zeamais (Coleoptera:
Curculionidae). Rev Colomb Entomol 34:
76 - 82.
Lorenzi AQ, Araujo DS, Kurtz BC. 2005.
Annonaceae das restingas do estado do Rio
de Janeiro. Rodriguésia 56: 85 - 96.
Maas PJM, Kamer HM, Junikka L, Mello-Silva R,
Hainer H. 2011a. Annonaceae of eastern and
south-eastern Brazil (Bahia, Espírito Santo,
Goiás, Minas Gerais, Mato Grosso, São
Paulo & Rio de Janeiro). Rodriguésia 52: 61
- 94.
Maas PJM, Westra LYT, Rainer H, Lobão AQ,
Erkens RAG. 2011b. An updated index to
genera, species, and infraspecific taxa of
Neotropical Annonaceae. Nord J Bot 29:
257 - 356.
Maas PJM, Lobão L, Rainer H. 2013. Annonaceae
na lista de Espécies da Flora do Brasil,
Jardim Botânico do Rio de Janeiro.
http://floradobrasil.jbrj.gov.br/jabot/flora
dobrasil/FB110219.
[Consulted November 12th, 2015].
Mikolajczak K, Madriga A, Rupprecht K, Hui YH,
Liu YM, Smith DL, McLaughlin JL. 1990.
Sylvaticin: a new cytotoxic and insecticidal
acetogenin from Rollinia sylvatica
(Annonaceae). Experientia 46: 324 - 327.
Miller JS. 2011. The discovery of medicines from
plants: a current biological perspective. Econ
Bot 65: 396 - 407.
Nelder JA, Wedderburn RWM. 1972. Generalized
linear models. J R Stat Soc 135: 370 - 384.
Okwulehie IC, Alfred NK. 2010. Fungi associated
with deterioration of sour-sop (Annona
muricata Linn.) fruits in Abia State, Nigeria.
Afr J Microbiol Res 4: 143 - 146.
R Development Core Team. 2012. R: A language and
environment for statistical computing. R
Foundation for Statistical Computing,
Vienna, Austria.
Rainer H. 2007. Monographic studies in the genus
Annona L. (Annonaceae): inclusion of the
genus Rollinia A.ST.-HIL. Ann Naturhist
Mus Wien - Serie B 108: 191 - 205.
Ramesha BT, Gertsch J, Ravikanth G, Priti V,
Ganeshaiah KN, Shahnker RU. 2011.
Biodiversity and chemodiversity: future
perspectives in bioprospecting. Curr Drug
Targets 12: 1515 - 1530.
Ribeiro LP. 2010. Bioprospecção de extratos
vegetais e sua interação com protetores de
grãos no controle de Sitophilus zeamais
Mots. (Coleoptera: Curculionidae),
Dissertação, Escola Superior de Agricultura
“Luiz de Queiroz”, Universidade de São
Paulo, Brasil.
Ribeiro LP, Vendramim JD, Bicalho KU, Andrade
MS, Fernandes JB, Moral RA, Demétrio
CGB. 2013. Annona mucosa Jacq.
(Annonaceae): a promising source of
bioactive compounds against Sitophilus
zeamais Mots. (Coleoptera: Curculionidae). J
Stored Prod Res 55: 6 - 14.
SAS Institute. 2011. Statistical Analysis System:
getting started with the SAS learning.
Version 9.2, Cary, NC.
Seffrin RC, Shikano I, Akhtar Y, Isman MB. 2010.
Effects of crude seed extracts of Annona
atemoya and Annona squamosa L. against the
cabbage looper Trichoplusia ni in the
laboratory and greenhouse. Crop Prot 29: 20
- 24.
Simões CMO, Schenkel EP. 2002. A pesquisa e a
produção brasileira de medicamentos a partir
de plantas medicinais: a necessária interação
da indústria com a academia. Rev Bras
Farmacogn 12: 35 - 40.
Sousa OV, Del-Vechio-Vieira G, Amaral MPH,
Pinho JJRG, Yamamoto CH, Alves MS.
2008. Antinociceptive and anti-inflammatory
effects of Duguetia lanceolata St.-Hil.
(Annonaceae) leaves ethanol extract. Lat Am
J Pharm 27: 398 - 402.
Tempone AG, Borborema SET, Andrade Jr HJ,
Gualda NCA, Yogi A, Carvalho CS,
Bachiega D, Lupo FN, Bonotto SV, Fischer
DCH. 2005. Antiprotozoal activity of
Brazilian plant extracts from isoquinoline
alkaloid-producing families. Phytomedicine
12: 382 - 390.
Throne JE, Weaver DK, Chew VB, James E. 1995.
Probit analysis of correlated data: multiple
observations over time at one concentration.
J Econ Entomol 88: 1510 - 1512.
Trindade RCP, Luna JD, Lima MRF, Da Silva PP,
Sant’ana AEG. 2011. Larvicidal activity and
Ribeiro et al.
Searching for sources of grain protectors in Neotropical Annonaceae
Boletín Latinoamericano y del Caribe de Plantas Medicinales y Aromáticas/232
seasonal variation of Annona muricata
(Annonaceae) extract on Plutella xylostella
(Lepidoptera: Plutellidae). Rev Colomb
Entomol 37: 223 - 227.
Valli M, Pivatto M, Danuello A, Castro-Gamboa I,
Silva DHS, Cavalheiro AJ, Araújo AR,
Furlan M, Lopes MN, Bolzani VS. 2012.
Tropical biodiversity: Has it been a potential
source of secondary metabolites useful for
medicinal chemistry? Quím Nova 35: 2278 -
2287.
Ware GW, Whitacre DM. 2004. Introducción a los
insecticidas. In: The Pesticide Book,
MeisterPro Information Resources, Ohio,
USA.
Wink M. 2003. Evolution of secondary metabolites
from an ecological and molecular
phylogenetic perspective. Phytochemistry
64: 3 - 9.
Zafra-Polo MC, Figadere B, Gallardo T, Tormo JR,
Cortez D. 1998. Natural acetogenins from
Annonaceae, synthesis and mechanisms of
action. Phytochemistry 48: 1087 - 1117.
... Chemical control is the most commonly used management strategy in the warehouse in Brazil and elsewhere (Braga et al., 2011;Ribeiro et al., 2016;Luz et al., 2017); however, the increase in insecticide use leads to greater selection pressure and an increase in the number of populations that are resistant to the insecticidal active ingredients, such as pyrethroids and organophosphates (Perry et al., 2011), a scenario that has stimulated the study and development of alternative strategies (Ribeiro et al., 2013). Additionally, the practice of applying many chemicals near consumption often poses a risk to workers and consumers due to potentially toxic residues on the grains (Stejskal et al., 2015). ...
... Pest management with botanical insecticides has attracted great social interest because they are generally safer and have greater selectivity than synthetic products (Isman, 2008;Isman, 2015;Ribeiro et al., 2016). Such products generally do not require highly qualified personnel, are inexpensive and have few environmental effects, and they can be produced at the same location as where the grains are stored (homemade preparations), making them easy to use (Ribeiro et al., 2014). ...
... Such products generally do not require highly qualified personnel, are inexpensive and have few environmental effects, and they can be produced at the same location as where the grains are stored (homemade preparations), making them easy to use (Ribeiro et al., 2014). Additionally, the difficulty of obtaining new synthetic molecules and the consequent high production costs have also stimulated studies with botanically derived insecticides (Ribeiro et al., 2013;Ribeiro et al., 2016;Singh, 2017). ...
Article
Full-text available
p> Background. Zabrotes subfasciatus (Boh., 1833) (Coleoptera: Chrysomelidae: Bruchinae), is considered one of the most important pest of stored beans. Objective. This study reports the possible toxicity and repellence of powders prepared from eight plant species against the Mexican bean weevil in two formulations (dry powder and sachets). Methodology. A 10 <sub>˟</sub> 2 factorial design (10 species <sub>˟</sub> 2 formulations) with 8 repetitions in a completely randomized design was employed. Pots with no powder were used as a negative control, and a pyrethroid insecticide [K-Obiol<sup>®</sup> 2 DP (deltamethrin, 0.5 g a.i. ton<sup>-1</sup>)] was applied as a positive control. Results. A mixture of powdered Chenopodium ambrosioides L., Ruta graveolens L. and Mentha pulegium L. added to bean grains was confirmed to be toxic to Z. subfasciatus adults with promising grain protector properties. C. ambrosioides powder had the same effect when in a sachet. A mixture of powdered R. graveolens , M. pulegium and C. ambrosioides with the beans inhibited weevil oviposition. The same effect was achieved for M. pulegium and C. ambrosioides in sachets. A mixture of powdered C. ambrosioides , M. pulegium , R. officinalis and R. graveolens repelled Z. subfasciatus adults from bean grains. Implications. This is the first report of using botanical derivatives by means of sachets or dry formulations, a pre-commercial purpose for aromatic plants with insecticidal/repellent activities. Conclusions. Sachets containing powdered C. ambrosioides and M. pulegium efficiently controlled the Mexican bean weevil in stored beans and constitute an useful tools for domestic grain stock or post-harvest management of organic grains.</p
... In addition to the economic impact of grain spoilers, the search for environmentally friendly strategies and low-risk compounds that can be included in integrated pest management (IPM) of stored products has been encouraged by concerns about pest resistance to currently available insecticides (Boyer et al., 2012;López-Castillo et al., 2018;Nayak et al., 2020), environmental risks involved in the routine use of xenobiotics, and detection of agrochemical residues in foodstuff (Teló et al., 2015;Han et al., 2016;Mebdoua and Ounane, 2019). Compounds originating from secondary plant metabolism (allelochemicals) are promising sources of new (bio)insecticides (Ribeiro et al., 2013(Ribeiro et al., , 2016(Ribeiro et al., , 2020 and have been widely studied to detect alternative tools that ensure effectiveness, safety, and selectivity standards (Viegas Jr, 2003;Ribeiro et al., 2016;Gonçalves et al., 2017), which are essential precepts for IPM programs. Such compounds are also an interesting option for developing high-value products within the bioeconomy approach (Valli et al., 2018), particularly in megadiverse countries such as Brazil (Abranches, 2020). ...
... In addition to the economic impact of grain spoilers, the search for environmentally friendly strategies and low-risk compounds that can be included in integrated pest management (IPM) of stored products has been encouraged by concerns about pest resistance to currently available insecticides (Boyer et al., 2012;López-Castillo et al., 2018;Nayak et al., 2020), environmental risks involved in the routine use of xenobiotics, and detection of agrochemical residues in foodstuff (Teló et al., 2015;Han et al., 2016;Mebdoua and Ounane, 2019). Compounds originating from secondary plant metabolism (allelochemicals) are promising sources of new (bio)insecticides (Ribeiro et al., 2013(Ribeiro et al., , 2016(Ribeiro et al., , 2020 and have been widely studied to detect alternative tools that ensure effectiveness, safety, and selectivity standards (Viegas Jr, 2003;Ribeiro et al., 2016;Gonçalves et al., 2017), which are essential precepts for IPM programs. Such compounds are also an interesting option for developing high-value products within the bioeconomy approach (Valli et al., 2018), particularly in megadiverse countries such as Brazil (Abranches, 2020). ...
... The number of emerged insects ranged from 333.4 to 366.0 compared to 602.4 emerged insects in the untreated control 60 . Ethanol seed extracts of Annona montana, Annona mucosa, Annona muricata and Annona sylvaticac completely inhabited the emergence of the insects 33 . In grains treated with ethanol leaf extracts from Annona mucosa, Annona montana, Duguetia lanceolata andAnnona muricata, the insects that emerged were 6.1, 8.4, 17.0 and 25.9 respectively compared to 51.4 in the negative control 33 . ...
... Ethanol seed extracts of Annona montana, Annona mucosa, Annona muricata and Annona sylvaticac completely inhabited the emergence of the insects 33 . In grains treated with ethanol leaf extracts from Annona mucosa, Annona montana, Duguetia lanceolata andAnnona muricata, the insects that emerged were 6.1, 8.4, 17.0 and 25.9 respectively compared to 51.4 in the negative control 33 . Methanol, hexane and hexane-methanol blend (1:1) extracts from leaves of Ocimum basilicum were evaluated for their adult emergence inhibition activity against S. zeamais at 25, 50, 75 and 100% concentrations. ...
Article
Full-text available
Insect pests cause significant losses to maize grains both in the fieldsand in storage. Maize weevil (Sitophilus zeamais) and larger grain borer (Prostephanus truncatus) are among the most destructive pests of maize in Africa. Currently, researchers have focused on identification of insecticidal plant extracts and compound through in-vivo and in-vitro experiments, as an alternative to synthetic chemicals which have been reported to have adverse effect on the environment.The aim of this study was to collate and review the fragmented information on plants extracts and compounds with insecticital activity against S. zeamais and P. truncatus and present recommendations for future research. Peer-reviewed articles were retrieved from Scopus, Science Direct, SciFinder and Google Scholar. This study led to identification of 123 plant species which have been examined for insecticidal activity focusing more on S. zeamais rather than P. truncatus.It is also evident that most studies end with the crude plant extracts. Effort towards identification of insecticidal principles from the plants is negligible despite the fact that insecticidal compounds from nature are preferred because they are environmentally safe. Future studies aimed in isolating and characterizing the active compounds from the plants is necessary. It is also necessary to develop plant based formulations to be used as altenatives in controlling the crop storage pests.
... C. B. Santos et al., 2019;Siqueira et al., 2001;Sousa et al., 2012;Z. W. Wang et al., 1988;Waterman, 1976) Studies demonstrating the insecticide and herbicide activities of the phenylpropanoids (Alves et al., 2020;Alves, Machado, Campos, Oliveira, & Carvalho, 2016;Gonçalves et al., 2017;Koona & Bouda, 2004;Koona & Bouda, 2006;Liu, Zhou, Liu, & Du, 2013;Popławski et al., 2000;Ribeiro et al., 2016;Siqueira et al., 2001;Vidotto et al., 2013;Z. W. Wang et al., 1988) Studies demonstrating the acaricide and larvicidal activity of the phenylpropanoids (Alves et al., 2015;Bhardwaj et al., 2010;de Sousa et al., 2020;Pares et al., 2021;A. ...
... These results were comparable to those obtained with a commercial insecticide based on deltamethrin (2 g.kg-1). A less pronounced effect against S. zeamais was verified by Ribeiro and co-workers who demonstrated that a treatment of corn grains with an ethanolic crude extract prepared from the leaves of D. lanceolata (3 g.kg-1) produced only 37.5% of mortality after 10 days (Ribeiro et al., 2016). Th toxic effect of two essential oils obtained from the stem barks and the aerial parts of D. furfuracea against Artemia saline, revealed that only the oil derived from the barks was active (LC50= 715.2 mg.mL -1 ) (Vidotto et al., 2013). ...
Article
Full-text available
The genus Duguetia encompasses 93 species, 63 of which are distributed in Brazil. However, only ten species had their chemical and biological profiles investigated so far. Although the alkaloids are the class of phytocompounds most studied, the interest in the phenylpropanoids is growing since distinct biological activities have been attributed to these derivatives lately. This review gathered studies describing the phytochemical distribution, methods of extraction, and biological activities of the phenylpropanoid from Duguetia species. It was evidenced that nonpolar solvents were able to provide the highest yield of 2,4,5-trimethoxystyrene, γ-asarone, and asaraldehyde which were mainly located at the ground stem barks. On the other hand, the α-asarone, elimicin, and (E)-methyl-isoeugenol were isolated from the polar extracts or the essential oils from the barks and leaves. The 2,4,5-trimethoxystyrene and asarone derivates were effective as crop protectors. The α- and γ-asarone, 2,4,5-trimethoxycinnamic acid, asaraldehyde, elemicin, and (E)-methyl-isoeugenol demonstrated anti-inflammatory, antidyslipidemic, antinociceptive, antibacterial, antidepressant, anxiolytic, insecticide, and larvicide activities. The Duguetia species can be a potential source of phenylpropanoids, therefore, future studies involving their extraction, identification, and application are still necessary and may culminate in the discovery of new drug candidates or natural agricultural defensives.
... This study is the first to report the insecticidal activity of the EO from X. brasiliensis against the FAW. Previous works conducted using the ethanol extract and soluble fraction in dichloromethane from the methanolic extract of X. brasiliensis did not find a pesticidal activity for Sitophilus zeamais (Motschulsky) (Coleoptera: Curculionidae) [67], Tetranychus tumidus Banks (Acari: Tetranychidae) [50] and S. frugiperda in an ingestion test [20]. Although there are few studies aimed at evaluating the chemical characterization of EOs from X. brasiliensis, it can be mentioned that spathulenol (40.8%) is the major component of the EO of the leaves of this plant [68]. ...
Article
Full-text available
The fall armyworm (FAW) Spodoptera frugiperda is a polyphagous pest that is difficult to control due to populations resistant to various active ingredients. Thus, the objective of this study was to evaluate the toxicity of essential oils (EOs) from the organs of Annona neolaurifolia, Duguetia lanceolata, and Xylopia brasiliensis, against the FAW and its natural enemy, Trichogramma pretiosum. The most active EOs were those from the leaves and stem bark of D. lanceolata, which presented LD90 to S. frugiperda equal to 70.76 and 127.14 µg of EO/larvae, respectively. The major compounds in the EO of D. lanceolata (leaves) were β-caryophyllene and caryophyllene oxide. Although individually inactive against the FAW, when combined, those compounds reduced the insect's probability of survival. However, the mortality was lower than that caused by EO. This result suggests that other components of EO contribute to the activity against FAW. Furthermore, the EO of the leaves from D. lanceolata presented low toxicity to the egg-larva stage of T. pretiosum, but was toxic to other phases. Thus, EO from D. lanceolata is potentially useful for developing new products to control S. frugiperda.
Article
The objective of this study was to evaluate the effects of growth-regulating insecticides of synthetic (e.g., Certero 480 SC, Intrepid 240 SC, Match EC and Mimic 240 SC) and botanical origins (e.g., Azamax 1.2 EC, Agroneem 850 EC, Azact 2.4 EC and Fitoneem 850 EC) on the biological parameters and fertility life table of Spodoptera frugiperda (J.E. Smith) under laboratory conditions. Larvae were fed insecticides that were incorporated into artificial diets. To develop the fertility life table, the following biological parameters were evaluated: survival at 7 days after infestation (d.a.i) and survivorship at adult eclosion, duration of the neonate-to-adult eclosion period, larval and pupal weights and total fecundity (number of total eggs per female). The results indicated that S. frugiperda neonates surviving LC 25 or LC 50 concentrations of the evaluated insecticides showed longer larval and egg-to-adult periods, lower larval and pupal weights and reduced fecundity, when compared to the control treatment. Larvae exposed to Azamax at LC 25 or LC 50 concentrations showed the greatest increase in generation duration (75 d). In addition, S. frugiperda adults emerged from pupae when larvae reared on an artificial diet containing growth regulating insecticides of synthetic and botanical origins produced fewer females per female per generation ( R o ). As well as, lower rates of natural population increase per day ( r m ) compared to insects fed the control diet. Our findings indicated that, neem-derived products and growth-regulating insecticides of synthetic origin may be employed within integrated management strategies that aim to keep populations of S. frugiperda below levels that cause economic damage. Similarly, they offer alternatives for insecticide resistance management programs.
Chapter
Full-text available
Objetivo: Analisar como a literatura tem abordado a utilização de extratos enquanto estratégia para o manejo de insetos na agricultura. Métodos: A metodologia partiu de uma abordagem qualitativa de caráter descritivo, estabelecida pela revisão de literatura. O Portal de periódicos da CAPES, a SciELO, a Elsevier e o Google Acadêmico foram os suportes utilizados para a identificação de artigos científicos, por meio dos descritores “Plant Extracts”, “Insect Management with Extract”. Resultados: As ações antrópicas, no caso da agricultura tradicional, são os principais responsáveis pelo desiquilíbrio no ecossistema, o que favorece o surgimento e/ou a proliferação de insetos-praga. No Brasil, as pragas que apresentam maior risco fitossanitário em culturas agrícolas são: Helicoverpa armígera; Chrysodexis includens; Spodoptera frugiperda; Anticarsia gemmatalis; Euchistus heros; Dichelops melacanthus; Bemisia tabasi; Antonomus grandis; Oryzophagus oryzae. Por esse motivo, diversas plantas têm sido estudadas e/ou utilizadas em cultivares, na forma de extrato ou óleo, devido às suas substâncias antimicrobianas, para o controle de patógenos na agricultura. Entre essas plantas estão o “nim”, a “citronela”, o “cravo-da-índia”, o “angico vermelho”, a “primavera vermelha” e o “tabaco”. Conclusão: É possível validar, de acordo com a literatura, a aplicabilidade dos extratos vegetais no manejo integrado de pragas, bem como de doenças causadas por patógenos.
Article
Full-text available
p> Background. The Annonaceae botanical family is a promising source of new insecticidal molecules that can be used to protect stored beans against Bruchinae beetles due to its variety of bioactive compounds such as alkaloids and acetogenins. Bruchinae beetles are major pests of stored beans in the world. These beetles feed on stored beans and they promote high levels of damages. Objective. The present study assessed the lethal and sublethal effects of crude ethanolic extracts prepared from different parts (leaves, branches and seeds) of Annona montana, Annona mucosa, Annona muricata and Annona reticulata against the Mexican bean weevil, Zabrotes subfasciatus (Boehman). Methodology. It was performed toxicological bioassays by spraying ethanolic extracts and fractions from Annona species on grains surface. The treated grains were infested with adults of Z. subfasciatus . It was evaluated the number of dead insects, eggs, F<sub>1</sub> progeny and damaged grains. Ethanolic extracts obtained by cold maceration [ratio 1:5 (v p<sup>-1</sup>)] were applied on grains surface (cv. Bolinha) at 1,500 ppm [extract (mg) grains (kg)<sup>-1</sup>]. Results. The seed extract from A. mucosa promoted 100% of mortality of Z. subfasciatus and completely inhibited the oviposition, the F<sub>1</sub> progeny emergence and the damage on grains. In addition, it was also estimated the LC<sub>50</sub> of the ethanolic extracts from seeds of Annona species. Interestingly, lethal concentrations varied according to Z. subfasciatus sex, and females supported higher doses than males. The ethanolic extract from seeds of A. mucosa presented the lowest LC<sub>50 </sub>value (571.82 mg kg<sup>-1</sup>) among all extracts. Thereby, it was submitted to liquid-liquid partition producing a hexane fraction and a remaining hydro-methanol phase. The former fraction killed 100% of insects whereas the later killed 20% of insects, but both of them promoted sublethal effects on Z. subfasciatus reducing the number of eggs and F<sub>1</sub> progeny. Implication. The ethanolic extracts of A. mucosa , A. montana , A. muricata and A. reticulata can be used as raw material for the formulation of a botanical insecticide that can help small farmers. Conclusion. The ethanolic extracts from seeds of A. mucosa , A. montana , A. muricata and A. reticulata are highly toxic to Zabrotes subfasciatus and protected the stored beans.</p
Article
The whitefly Bemisia tabaci(Gennadius) MEAM 1 is one of the main insect species that colonize tomato plants and cause direct and indirect damage. The use of botanical derivatives may be a valuable method of insect control to reduce the inappropriate use of synthetic insecticides on crops. In this study, we evaluated the bioactivity of ethanolic extracts prepared from Annonaceae species compared to that of the commercial insecticides based on acetogenins (Anosom® 1 EC, anonine 10,000 mg L-1) and thiamethoxam (Actara® 250 WG) on eggs, nymphs, and adults of the whitefly in tomato. Initially, the effects of the ethanolic seed extracts of Annona mucosa (Jacq.), Annona muricata L., and Annona sylvatica A.St.-Hil on adult insect behavior were evaluated. The rates of infestation and oviposition deterrence indicated the inhibitory effects of the extract of A. muricata (500 mg L-1). Then, the possible systemic effects of the extracts were evaluated; however, no effects on nymphal development or insect viability were observed. The LC50 and LC90 of the ethanolic extract of A. mucosa seeds at 500 mg L-1 (10.83 and 200.24 mg L-1, respectively) were estimated and were used in ovicidal tests and compared to positive (Actara® 250 WG and Anosom® 1 EC), and negative controls (water: acetone, 1:1 v/v). At LC90, fewer eggs (35.00%) had hatched at 13 days after application than in the other treatments. The results of this study demonstrate the potential use of botanical derivatives of Annona spp. for the management of B. tabaci MEAM 1 in tomato.
Article
The poultry red mite Dermanyssus gallinae (De Geer) is the most important haematophagous ectoparasite in the poultry industry. The use of synthetic acaricides for this control is presenting risks related to human food. In this sense, plant secondary metabolites are promising for controlling this pest. Thus, this study aimed to evaluate the acaricidal activity of Duguetia lanceolata A.St.-Hil. (stem bark), Xylopia emarginata Mart. (stem bark), and Xylopia sericea A.St.-Hil. (stem bark and fruits) against D. gallinae. Additionally, the secondary metabolite profile of the X. emarginata was analysed by UFLC-DAD-ESI(+)-MS/MS (micrOTOF-QII) and data analysis was performed using the Molecular Networking. In a topical application test, all plant species tested showed bioactivity, in that order of toxicity with the respective probability survival: X. emarginata (stem bark) (0.28) > X. sericea (stem barks) (0.35) > X. sericea (fruits) and D. lanceolata (stem bark) (0.47). The most promising results were found for X. emarginata (LC50 = 331.769 μg/cm2). It is noteworthy that the LC50 of the insecticide cypermethrin was 1234.4 μg/cm2, which was 73% higher than that of X. emarginata. The metabolomic profile of X. emarginata revealed the presence of alkaloids, amides, terpenoids, and phenolic compounds. This is the first report of X. emarginata acaricidal activity against D. gallinae and exploratory chemical analysis by untargeted metabolomics and the molecular network of this plant.
Article
Full-text available
The family Annonaceae is represented in the restingas (sandy coastal plains) of Rio de Janeiro State by nine species: Anaxagorea dolichocarpa Sprague & Sandwith, Annona acutiflora Mart., A. glabra L., A. montana Macfad., Duguetia sessilis (Vell.) Maas, Guatteria nigrescens Mart., Oxandra nitida R.E.Fr., Xylopia ochrantha Mart. and X. sericea A.St.-Hil. A species key, short descriptions, illustrations and comments on the phenology, geographic distribution, habitats and uses are included.
Article
Full-text available
Potato psyllid, Bactericera cockerelli (Sulcen), nymphs were treated with extracts of Annona muricata L., Carica papaya L., Euphorbia dentate Michx, Thuja occidentalis L., Sapindus saponaria L., and Azadirachta indica A. Juss. (neem) as a commercial check in the laboratory. At 72 hours, most potato psyllid nymphs died (98 and 100% mortalities) from A. muricata extract from seeds, at concentrations of 2,500 and 5,000 ppm, respectively, followed by A. indica oil that caused 91 and 100% mortality at concentrations of 2,000 and 2,500 ppm, respectively. A. muricata seed extract was the most effective insecticide in the study.
Article
Full-text available
New control methods are necessary for stored grain pest management programs due to both the widespread problems of insecticide-resistance populations and the increasing concerns of consumers regarding pesticide residues in food products. Thus, this study evaluated the bioactivity of extracts and fractions obtained from different structures (leaves, branches, and seeds) of Annona mucosa (Annonaceae) against Sitophilus zeamais (Coleoptera: Curculionidae), which is a primary insect pest of stored cereals in tropical conditions. In the screening assay, the most promising treatments were extracts prepared from the seeds of Annona mucosa in hexane and dichloromethane (LC90 values of 259.31 and 425.15 mg kg−1, respectively) and, to a lesser extent, an extract prepared from the leaves in hexane (LC90 of 1047.15 mg kg−1). Based on these results and the chromatographic profile of the bioactive crude extracts, the extract prepared from the seeds in hexane was fractionated by liquid–liquid partitioning. The dichloromethane and hydroalcoholic fractions exhibited insecticidal activity against S. zeamais, and no significant difference was observed between these two fractions. The chemical analyses (1H NMR, HPLC, and TLC) showed the presence of alkaloids and acetogenins in the bioactive fractions, which are likely related to the observed bioactivity. Thus, A. mucosa, particularly its seeds, is a promising source of compounds that can be used as a prototype model and/or a biorational insecticide for the control of S. zeamais in stored cereals.
Article
Duguetia lanceolata St.-Hil. (Annonaceae), popular known as pindaíba, is used in folk medicine as anti-inflammatory, cicatrizing and antimicrobial. The aim of the present study was to investigate the analgesic and anti-inflammatory effects of D. lanceolata ethanol extract by acetic acid writhing, paw licking induced by formalin, hot plate and paw edema tests. The ethanol extract reduced (p < 0.05) the abdominal contortions (10 mg/kg = 58.37 ± 1.90; 50mg/kg = 54.37 ± 2.14; 100 mg/kg 41.50 ± 2.04 and 200 mg/ kg = 34.25 ± 1.76) when compared with control group. Both phases of paw lick were inhibited (p < 0.05): 1st phase (50 mg/kg = 72.50 ± 1.97; 100 mg/kg = 65.87 ± 2.27 and 200 mg/kg = 47.75 ± 2,20 and 2nd phase (50 mg/kg = 71.87 ± 2.78; 100 mg/kg = 64.25 ± 1.72 and 200 mg/kg = 56.62 ± 2.32). After 30 min of treatment, the reaction time on the hot plate increased since dose 50 mg/kg (7.62 ± 0.50) with maximal effect at dose 200 mg/kg (12.87 ± 0.87). The doses of 50, 100 and 200 mg/kg had significant effect to reduce the paw edema. These results suggest that D. lanceolata could constitute a source of active substances with antinociceptive and anti-inflammatory effects.
Article
The technique of iterative weighted linear regression can be used to obtain maximum likelihood estimates of the parameters with observations distributed according to some exponential family and systematic effects that can be made linear by a suitable transformation. A generalization of the analysis of variance is given for these models using log- likelihoods. These generalized linear models are illustrated by examples relating to four distributions; the Normal, Binomial (probit analysis, etc.), Poisson (contingency tables) and gamma (variance components). The implications of the approach in designing statistics courses are discussed.
Article
The in-vitro antifungal, antibacterial, insecticidal, phytotoxic and cytotoxic properties of A. squamosa (Annonaceae) were studied. Ethanolic and pet. ether extracts of fruit pulp and seeds were tested for antifungal activity in-vitro against six fungi, Trichophyton longifusus, Candida albicans, Aspergillus flavus, Microsporium canis, Fusarium solani,Candida glaberata. besides this it is also tested for antibacterial activity against six bacterianamely: Escherachia coli. Bacillus subtilis. Shigella flexeneri, Staphylococcus aureus, Pseudomonas aeruginosa and Salmonella typhi. The degree of susceptibility varied with different solvents. The alcoholic extract of seeds showed higher susceptibility for Trichophyton longifusus. Bioassays of the crude alcoholic and pet. ether extracts, in the brine shrimp test for cytotoxicity, evaluated relative potencies. Alcoholic extract of seeds was 1000 times more potent than that of etoposide. The pet. ether extract of seeds of A. squamosa exhibited 57.8% phytogrowth inhibitory activity on seedlings of Lamna minor.It also exhibited the 50% growth inhibition of insects.
Article
Anona muricata commonly known as sour-sop, belong to the family Annonaceae. It is a small slender tropical tree usually grown for its large fleshy and juicy fruits. The fruit of A. muricata plays an important role in the diet of many in many parts of the tropics including Nigerian. Unfortunately, the usefulness of the fruits of Anona is decimated by many fungi species. Investigation of the fungi that cause the deterioration of sour-sop (Anona muricata) was carried out in order to recommend the appropriate control measures. Mature fruits of A. muricata were collected from different locations in Abia State. Isolation, characterization and identification of fungi were made by plating washings from skin surface and extracted juice from pulp of the fruits in test tubes containing potato dextrose agar (PDA) and yeast malt extract agar (Difco) into which streptomycin sulphate was incorporated to inhibit bacterial growth. Inoculated tubes were incubated at 28 ± 2°C for 48 h. Control experiment was carried out with sterile peptone water instead of using the washings. The following filamentous and yeast fungi were isolated from the skin surface and pulp of the fruits: Aspergillus flavus, Aspergillus niger, Botryodiplodia theobromae, Colletotrichum sp., Fusarium solani, Mucor sp., Penicillium chrysogenium, Penicillium sp., Rhizopus stolonifer and Rigidoporus sp., Candida albicans, Saccharomyces cerevisiae, Saccharomyces rouxii and Torulopsis spp. The pathogenicity tests of the filamentous fungi isolated were carried out on mature green Anona fruits and were found to be pathogenic. The filamentous fungi were mainly responsible for the deterioration of the fruits of A. muricata in Abia state while the yeasts were fermentative.